179
Views
0
CrossRef citations to date
0
Altmetric
Research Article

Bridgeland stability conditions and skew lines on ℙ3

&
Pages 3081-3114 | Received 11 Dec 2022, Accepted 08 Jan 2024, Published online: 19 Feb 2024

Abstract

Inspired by Schmidt’s work on twisted cubics, we study wall crossings in Bridgeland stability, starting with the Hilbert scheme Hilb2m+2(P3) parametrizing pairs of skew lines and plane conics union a point. We find two walls. Each wall crossing corresponds to a contraction of a divisor in the moduli space and the contracted space remains smooth. Building on work by Chen–Coskun–Nollet, we moreover prove that the contractions are K-negative extremal in the sense of Mori theory and so the moduli spaces are projective.

2020 MATHEMATICS SUBJECT CLASSIFICATION:

1 Introduction

After the introduction of Bridgeland’s manifold of stability conditions on a triangulated category [Citation6], several applications to the study of the birational geometry of moduli spaces have appeared: the moduli space is viewed as parameterizing stable objects in the derived category of some underlying variety X, and the question is how the moduli space changes as the stability condition varies. This is the topic of wall crossing in the stability manifold. We refer to [Citation15] for an overview and in particular for examples of the success of this viewpoint in cases where X is a surface. For threefolds and notably in the case X=P3, important progress was made by Schmidt [Citation20], allowing among other things a study of wall crossings for the Hilbert scheme of twisted cubics (see also Xia [Citation22] for further work on this case; additional examples in the same spirit have been investigated by Gallardo–Lozano Huerta–Schmidt [Citation11] and Rezaee [Citation19]). The case considered in the present text is that of pairs of skew lines in P3 and their deformations. This is analogous to twisted cubics in the sense that a twisted cubic degenerates to a plane nodal curve with an embedded point much as a pair of skew lines degenerates to a pair of lines in a plane together with an embedded point.

More precisely we study wall crossing for the Hilbert scheme Hilb2m+2(P3) of subschemes YP3 with Hilbert polynomial 2m+2. It has two smooth components C and S: a general point in C is a conic-union-a-point Y=C{P} and a general point in S is a pair of skew lines Y=L1L2. Note that when a line pair is deformed until the two lines meet, the result is a pair of intersecting lines with an embedded point at the intersection, and this can also be viewed as a degenerate case of a conic union a point.

For an appropriately chosen Bridgeland stability condition on the bounded derived category of coherent sheaves Db(P3), the ideal sheaves IY can be viewed as the stable objects with fixed Chern character, say v=ch(IY). When deforming the stability condition, we identify two walls, separating three chambers. Getting slightly ahead of ourselves, the situation is illustrated in in Section 3.2: α and β are parameters for the stability conditions considered and we restrict ourselves to the region to the immediate left in the picture of the hyperbola β2α2=4 (the role of this boundary curve is explained in Section 2.4). In this region we have the two walls W1 and W2 separating three chambers, labeled by Roman numerals as in the figure. Let MIMII, and MIII be the moduli spaces of Bridgeland stable objects with Chern character v in each chamber, considered as algebraic spaces (for existence see Piyaratne–Toda [Citation18, Corollary 4.23]).

Our first main result contains the set-theoretical description of these moduli spaces:

Theorem 1.1.

  1. MI is Hilb2m+2(P3) with its two components S and C described above.

  2. MII consists of:

    1. Ideal sheaves IY for YHilb2m+2(P3) not contained in a plane.

    2. Non-split extensions FP,V in

      0IP/V(2)FP,VOP3(1)0 (1.1)

      for VP3 a plane and PV. Moreover, FP,V is uniquely determined up to isomorphism by the pair (P, V).

  3. MIII consists of:

    1. Ideal sheaves IY for YHilb2m+2(P3) a pair of disjoint lines or a pure double line.

    2. Non-split extensions GP,V in

      0OV(2)GP,VIP(1)0 (1.2)

      with P and V as above. Moreover, GP,V is uniquely determined by the pair (P, V).

To locate walls and classify stable objects we employ the method due to Schmidt [Citation20], which involves “lifting” walls from an intermediate notion of tilt stability. Schmidt considers as an application the Hilbert scheme Hilb3m+1(P3): it parameterizes twisted cubics and plane cubics union a point. This was our starting point and we can apply many of Schmidt’s results directly, although modified or new arguments are needed as well. The end result is closely analogous in the two cases, with two wall crossings of the same nature. In the twisted cubic situation, however, Schmidt also finds an additional “final wall crossing” where all objects are destabilized. This has no direct analogy in our case.

Next we describe the moduli spaces geometrically, guided by the set theoretic classification of objects above; this leads to contractions of the two smooth components C and S of MI=Hilb2m+2(P3). First introduce notation for the loci that are destabilized by the two wall crossings according to the above classification:

Notation 1.2.

  1. Let EC be the divisor consisting of all planar YC.

  2. Let FS be the divisor consisting of all YS having an embedded point.

Thus the locus (II)(i) is (CE)S and the locus (III)(i) is SF. On the other hand both loci (II)(ii) and (III)(ii) are parameterized by the incidence variety IP3×Pˇ3 consisting of pairs (P, V) of points PP3 inside a plane VPˇ3. The process of replacing E and F by I can be realized as contractions of algebraic spaces: E and F may be viewed as projective bundles over I, and in Section 4 we apply Artin’s contractibility criterion to obtain smooth algebraic spaces C and S each containing the incidence variety I as a closed subspace, and birational morphisms ϕ:CCψ:SSwhich are isomorphisms outside of E, respectively F, and restrict to the natural maps EI, respectively FI. Moreover EC is disjoint from S, so the union CS makes sense as the gluing together of (CE)S and C. We can then state our second main result:

Theorem 1.3.

  1. MII is isomorphic to CS.

  2. MIII is isomorphic to S.

To prove the theorem it suffices to treat the contracted spaces as algebraic spaces. However, they turn out to be projective varieties: the contractions are in fact K-negative extremal contractions in the sense of Mori theory. The case of SS can be found in previous work by Chen–Coskun–Nollet [Citation8] and in fact it turns out that S is a Grassmannian; see Section 4.2. Inspired by this work, we exhibit in Section 5 the map CC as a K-negative extremal contraction. This may be contrasted with Schmidt’s approach in the twisted cubic situation [Citation20], where projectivity of the moduli spaces is proved by viewing them as moduli of quiver representations.

In Section 2, we list the background results that we need, in particular, we briefly recall the construction of stability conditions on threefolds, along with the notion of tilt-stability. In Section 3 we apply Schmidt’s machinery to prove Theorem 1.1. In Section 4 we study universal families and prove Theorem 1.3. Finally, in Section 5 we work out the Mori cone of C.

We work over C. Throughout and in particular in Section 4, intersections and unions of subschemes are defined by the sum and intersection of ideals, respectively, and inclusions and equalities between subschemes are meant in the scheme theoretic sense. The relative ideal of an inclusion ZY of two closed subschemes of some ambient scheme is the ideal IZ/YOY defining Z as a subscheme of Y.

2 Preliminaries

After detailing the two components of Hilb2m+2(P3), we collect notions and results from the literature surrounding Bridgeland stability and wall crossings for smooth projective threefolds. There are no original results in this section.

2.1 The Hilbert scheme and its two components

It is known that Hilb2m+2(P3) has two smooth components C and S, whose general points are conics union a point and pairs of skew lines, respectively. A quick parameter count yields dimC=11 and dimS=8. We refer to Lee [Citation13] for an overview, to Chen–Nollet [Citation9] for the smoothness of C and to Chen–Coskun–Nollet [Citation8] for the smoothness of S. In fact, the referenced works show that C is the blowup CP3×Hilb2m+1(P3)

along the universal conic ZP3×Hilb2m+1(P3) and S is the blowup SSym2(G(2,4))along the diagonal in the symmetric square of the Grassmannian G(2, 4) of lines in P3. In other words, it is the Hilbert scheme Hilb2(G(2,4)) of finite subschemes in G(2, 4) of length two.

We pause to make some preparations regarding embedded points: by a curve CP3 with an embedded point at PC we mean a subscheme YP3 such that CY and the relative ideal IC/Y is isomorphic to k(P). This makes sense even when we allow C to be singular or nonreduced. Embedded points are in bijection with normal directions, i.e. lines in the normal space to CP3 at P: explicitly in our situation, suppose that P=(0,0,0) in local affine coordinates x, y, z, and C is a conic defined by the ideal IC=(q(x,y),z), where q is a quadric vanishing at the origin. A normal vector to C may be viewed as a k-linear map ϕ:IC/mPICkgiven by say a=ϕ(q) and b=ϕ(z). Thus (a:b)P1 parameterizes normal directions and the corresponding scheme Y with an embedded point at P is defined by the kernel of the induced map ICk, which is IY=mPIC+(bqaz).

Example 2.1.

With notation as above, consider the degenerate conic defined by q(x,y)=xy. When (a:b)=(1:0) we obtain the planar scheme Y in the xy-plane given by IY=(x2y,xy2,z)=(xy,z)(z,x2,y2)where the final form exhibits Y as the scheme theoretic union of a pair of lines and a thickening of the origin in the xy-plane. At the other extreme (a:b)=(0:1) we obtain IY=(zy,zx,yz,z2)=(xy,z)(x,y,z)2 which we label spatial: the subscheme YP3 is not contained in any nonsingular surface since it contains a full first order infinitesimal neighborhood around P. It is easy to check that for the remaining values of (a:b) the corresponding scheme Y is neither planar nor spatial. Analogous observations hold for the double line q(x,y)=y2.

With these preparations, we next list all elements YHilb2m+2(P3), including degenerate cases. We follow Lee [Citation13, Section 3.5], to which we refer for further details and proof that the following list is exhaustive.

The component C parameterizes subschemes Y of the following form: let C be a conic in a plane VP3, possibly a union of two lines or a planar double line. Then Y is either the disjoint union of C and a point PP3, or C with an embedded point at PC. If C is nonsingular at P, embedded points correspond to normal directions, parameterized by a P1. Since even degenerate conics are complete intersections, also embedded point structures at a singular or nonreduced point P form a P1 as in Example 2.1, and among these exactly one is planar (Y contained in a plane) and exactly one is spatial (Y contains the first order infinitesimal neighborhood of P in P3).

The component S parameterizes pairs Y=L1L2 of skew lines, together with its degenerations. These are of the following three types: (1) a pair of incident lines L1L2 with a spatial embedded point at the intersection point, (2) a planar double line with a spatial embedded point, or (3) a double line in a quadric surface, i.e. there is a line L in a nonsingular quadric surface Q such that Y is the effective divisor 2LQ, but viewed as a subscheme of P3. We label this case the pure double line, where purity refers to the lack of embedded components.

Clearly, then, CS consists of the incident or planar double lines with a spatial embedded point.

2.2 Stability conditions and walls

Let X be a smooth projective threefold over C and fix a finite rank lattice Λ equipped with a homomorphism K(X)Λ from the Grothendieck group of coherent sheaves modulo short exact sequences. On P3 we will take Λ=ZZ12Z16Z equipped with the Chern character map ch:K(P3)Λ.

Recall [Citation4–6] that a Bridgeland stability condition σ=(A,Z) on X (with respect to Λ) consists of

  1. an abelian subcategory ADb(X), which is the heart of a bounded t-structure, and

  2. a stability function Z, which is a group homomorphism

    Z:ΛC

    whose value on any nonzero object EA is in the upper half plane

    H={zC|Iz0}R0.

This is subject to a list of axioms which we will not give (see [Citation4, Section 8]). We may then partially order the nonzero objects in A by their slope λσ=R(Z)/I(Z)R{+}.

This yields a notion of σ-stability and σ-semistability for objects in A in the usual way by comparing the slope of an object with that of its sub- or quotient objects. These notions extend to Db(X) by shifting in the sense of triangulated categories.

Bridgeland’s result [Citation6, Theorem 1.2] gives the set StabΛ(X) of stability conditions the structure of a complex manifold, and for a given uΛ, it admits a wall and chamber structure: if an object E goes from being stable to being unstable as the stability condition σ varies, σ has to pass through a stability condition for which E is strictly semistable. This observation leads to the definition of a wall:

Definition 2.2.

Fix uΛ.

  1. Numerical walls: A numerical wall is a nontrivial proper solution set

    Wv,w={σStabΛ(X)|λσ(v)=λσ(w)}

    where u=v+w in Λ.

  2. Actual walls: Let Wv,w be a numerical wall, defined by classes v, w satisfying u=v+w. A subset VWv,w is an actual wall if for each point σV, there is a short exact sequence

    0FEG0

    in A of σ-semistable objects E,F,G with classes u, v, w in Λ, respectively, such that λσ(F)=λσ(E)=λσ(G).

    When the context is clear, we drop the word “actual” and just say “wall”.

Given a union of walls, we refer to each connected component of its complement in StabΛ(X) as a chamber. By the arguments in [Citation7, Section 9] there is a locally finite collection of (actual) walls in StabΛ(X), each being a closed codimension one manifold with boundary, such that the set of stable objects in A with class uΛ remains constant within each chamber, and there are no strictly semistable objects in a chamber.

We say that a short exact sequence as in (ii) above defines the wall. Relaxing this, an unordered pair {F,G} defines the wall if there is a short exact sequence in either direction (i.e. we allow the roles of sub and quotient objects to be swapped) as in (ii). Semistability of F and G is automatic, i.e. it follows from semistability of E and the equality between slopes, but see Remark 2.4.

Remark 2.3.

A very weak stability condition (A,Z) is a weakening of the above concept (see Piyaratne–Toda [Citation18]) where Z is allowed to map nonzero objects in A to zero. One may define an associated slope function λ as before, with the convention that λ(E)=+ also when Z(E)=0. An object EA is declared to be stable if every nontrivial subobject FE satisfies λ(F)<λ(E/F), and semistable when nonstrict inequality is allowed. With this definition one avoids the need to treat cases where Z(F)=0 or Z(E/F)=0 separately. We will not need to go into detail.

2.3 Construction of stability conditions on threefolds

We next recall the “double tilt” construction of stability conditions by Bayer–Macrì–Toda [Citation5]. For this it is necessary to assume that the threefold X satisfies a certain “Bogomolov inequality” type condition [Citation4, Conjecture 4.1]), which is known in several cases including P3 [Citation14]. Fix a polarization H on X; on P3 this will be a (hyper)plane.

2.3.1 Slope stability

Let βR. The twisted Chern character of a sheaf or a complex E on X is defined by chβ(E)=eβHch(E). Its homogeneous components are ch0β(E)=ch0(E)=rk(E)ch1β(E)=ch1(E)βHch0(E)ch2β(E)=ch2(E)βHch1(E)+β2H22ch0(E)ch3β(E)=ch3(E)βHch2(E)+β2H22ch1(E)β3H36ch0(E).

The twisted slope stability function on the abelian category Coh(X) of coherent sheaves is given by (2.1) μβ(E)={H2ch1β(E)H3ch0β(E)if rk(E)0,+else.(2.1)

This is the slope of a very weak stability condition. Notice that μβ(E)=μ(E)β, where μ(E) is the classical slope stability function. A sheaf ECoh(X) which is (semi)stable with respect to this very weak stability condition is called μβ-(semi)stable (or slope (semi)stable).

2.3.2 Tilt stability

Next, define the following full subcategories of Coh(X) Tβ={ECoh(X)|  Any quotient EG satisfies μβ(G)>0},Tβ={ECoh(X)| Any subsheaf FE satisfies μβ(F)0}.

The pair (Tβ,Tβ) is a torsion pair [Citation7, Definition 3.2] in Coh(X). Tilt the category Coh(X) with respect to this torsion pair and denote the obtained heart by Cohβ(X)=Tβ[1],Tβ. Thus every object ECohβ(X) fits in a short exact sequence (2.2) 0H1(E)[1]EH0(E)0(2.2) with H1(E)Tβ and H0(E)Tβ.

Let (α,β)R>0×R, and let (2.3) Zα,βtilt(E)=Hch2β(E)+α22H3ch0β(E)+iH2ch1β(E).(2.3)

The associated slope function is να,β(E)={Hch2β(E)α22H3ch0β(E)H2ch1β(E)if H2ch1β(E)0,+else,

(see [Citation15, Section 9.1] for more details).

By [Proposition B.2 (case B=βH)], the pair (Cohβ(X),Zα,βtilt) is a very weak stability condition continuously parameterized by (α,β)R>0×R. An object ECohβ(X) which is (semi)stable with respect to this very weak stability condition is called να,β-(semi)stable (or tilt (semi)stable). Moreover the parameter space R>0×R admits a wall and chamber structure [Citation4, Proposition B.5], in which walls are nested semicircles centered on the β-axis, or vertical lines (we view α as the vertical axis) [Citation20, Theorem 3.3], and a numerical wall is either an actual wall everywhere or nowhere. We refer to them as “tilt-stability walls” or “ν-walls” interchangeably.

Remark 2.4.

Walls in the parameter space for tilt stability are defined analogously to Definition 2.2. With tilt-stability in mind we make the following observation: let E be strictly semistable with respect to a very weak stability condition (A,Z). By definition this means that E is semistable and there exists a short exact sequence 0FEG0in A such that all three objects share the same slope. This implies that F is semistable, since any destabilizing subobject of F would also destabilize E. When very weak stability conditions are allowed, however, G may not be semistable: this happens exactly when G has finite slope and there is a nontrivial subobject GG such that Z(G)=0. In this case let GG be the maximal such subobject: then G/G is semistable (it is in fact the final factor in the Harder–Narasimhan filtration of G) and has the same slope as G. Moreover the kernel F of the composite EGG/G has the same slope as F. Thus in the short exact sequence 0FEG/G0

all objects are semistable and of the same slope. So when looking for walls we may as in Definition 2.2 assume all objects in the defining short exact sequence to be semistable, even in the very weak situation.

The following is the Bogomolov inequality for tilt-stability:

Proposition 2.5.

[Citation5, Corollary 7.3.2] Any να,β-semistable object ECohβ(X) satisfies Δ¯H(E)=(H2ch1β(E))22H3ch0β(E)Hch2β(E)0.

2.3.3 Bridgeland stability

Define the following full subcategories of Cohβ(X): Tα,β={ECohβ(X)| Any quotient EG satisfies να,β(G)>0},Tα,β={ECohβ(X)|  Any subsheaf FE satisfies να,β(F)0}.

They form a torsion pair. Tilting Cohβ(X) with respect to this pair yields stability conditions (Aα,β(X),Zα,β,s) ([Citation4, Theorem 8.6, Lemma 8.8]) on X, where Aα,β(X)=Tα,β[1],Tα,β and (2.4) Zα,β,s=ch3β+α2(16+s)H2ch1β+i(Hch2βα22H3ch0β).(2.4)

The slope function of Zα,β,s is given by λα,β,s(E)=ch3β(E)α2(16+s)H2ch1β(E)Hch2β(E)α22H3ch0β(E)with λα,β,s(E)=+ when Hch2β(E)=α22H3ch0β(E).

An object EAα,β(X) which is (semi)stable with respect to this stability condition is called λα,β,s-(semi)stable.

The stability conditions (Aα,β(X),Zα,β,s) are continuously parameterized by (α,β,s)R>0×R×R>0 [Citation4, Proposition 8.10]. We refer to walls in R>0×R×R>0 as “λ-walls”.

The following lemma allows us to identify moduli spaces of slope-stable sheaves with moduli spaces of tilt-stable sheaves, given the right conditions:

Lemma 2.6.

[Citation11, Lemma 1.4] On P3, let v=(v0,v1,v2,v3)Λ satisfy μβ(v)>0 and assume (v0, v1) is primitive. Then an object ECohβ(P3) with ch(E)=v is να,β-stable for all α0 if and only if E is a slope stable sheaf.

2.4 Comparison between ν-stability and λ-stability—after Schmidt

Let E be an object in Db(X). Throughout this section, let (α0,β0)R>0×R satisfy να0,β0(E)=0, and fix s > 0. We shall summarize a series of results by Schmidt [Citation20] enabling us to compare walls and chambers with respect to ν-stability with those of λ-stability. (Looking ahead to our application illustrated in and , the dashed hyperbola is the solution set to να,β(E)=0.)

Fig. 1 The unique semicircular wall W (solid) for ν-stability and the hyperbola (dashed) from EquationEquation (3.3).

Fig. 1 The unique semicircular wall W (solid) for ν-stability and the hyperbola (dashed) from EquationEquation (3.3)(3.3) β2−α2=4.(3.3) .

Fig. 2 The two walls W1 and W2 (solid) for λ-stability separating three chambers, together with the hyperbola (dashed) from EquationEquation (3.3).

Fig. 2 The two walls W1 and W2 (solid) for λ-stability separating three chambers, together with the hyperbola (dashed) from EquationEquation (3.3)(3.3) β2−α2=4.(3.3) .

Consider the following conditions on E:

  1. E is a να0,β0-stable object in Cohβ0(X).

  2. E is a λα,β,s-stable object in Aα,β(X), for all (α,β) in an open neighborhood of (α0,β0) with να,β(E)>0.

  3. E is a λα,β,s-semistable object in Aα,β(X), for all (α,β) in an open neighborhood of (α0,β0) with να,β(E)>0.

  4. E is a να0,β0-semistable object in Cohβ0(X).

Obviously there are implications (1) (4) and (2) (3). The following says that, under a mild condition on ch(E), there are in fact implications (1)(2)(3)(4)so that λ-stability in a certain sense refines ν-stability:

Theorem 2.7

(Schmidt). The implication (1)(2) above always holds. If H2ch1β0(E)>0 then also the implication (3)(4) holds.

For the proof we refer to Schmidt [Citation20]: the first implication follows from Lemma 6.2 in loc. cit. and the second follows from Lemmas 6.3 and 6.4 (in the reference the additional condition Δ¯H(E)>0 appears; this is harmless but redundant as it is not used in the proof).

Schmidt furthermore compares walls for ν-stability and λ-stability, for objects E in some fixed class vΛ. Let (2.5) FEGF[1](2.5) be a triangle in Db(X) with E in class v.

  • Say that (2.5) defines a ν-wall through (α0,β0) if FEG are να0,β0-semistable objects in Cohβ0(X) and να0,β0(F)=να0,β0(G) (which is thus zero).

  • Say that (2.5) defines a λ-wall at the ν-positive side of (α0,β0) if there is an open neighborhood U of (α0,β0) such that, writing W={(α,β)U|να,β(v)>0 and λα,β,s(F)=λα,β,s(G)}

the following holds: (α0,β0) is in the closure of W and F,E,G are λα,β,s-semistable objects in Aα,β(X) for all (α,β)W.

Note that the assumption that F,E,G are all in Cohβ0(X) or in Aα,β(X) implies that the triangle (2.5) is a short exact sequence 0FEG0in that abelian category.

Theorem 2.8

(Schmidt). Let E be an object in Db(X) and let (α0,β0)R>0×R such that να0,β0(E)=0 and H2ch1β0(E)>0.

  1. If a triangle (2.5) defines a λ-wall on the ν-positive side of (α0,β0), then it also defines a ν-wall through (α0,β0).

  2. Suppose a triangle (2.5) defines a ν-wall through (α0,β0) and FG are να0,β0-stable. Moreover let

    W={(α,β)|να,β(v)>0 and λα,β,s(F)=λα,β,s(G)}

    and suppose there are points (α,β)W arbitrarily close to (α0,β0) such that να,β(F)>0 and να,β(G)>0. Then (2.5) defines a λ-wall on the ν-positive side of (α0,β0), namely W.

For the proof we refer to Schmidt [Citation20]: part (1) is Schmidt’s theorem 6.1(1) and part (2) is the special case n = 1 of Schmidt’s theorem 6.1(4). To align the notation, in part (1) Schmidt’s F,E,G are our F[1],E[1],G[1]. To apply theorem 6.1(1) these are required to be λα0,β0,s-semistable objects in Aα0,β0(X); this is ensured by Schmidt’s lemma 6.3.

To control how the set of stable objects changes as a λ-wall is crossed, we take advantage of the fact that the λ-walls we obtain are defined by short exact sequences with stable sub- and quotient objects (in other words, only two Jordan–Hölder factors on the wall) and apply:

Proposition 2.9.

Suppose F and G are λα,β,s-stable objects in Aα,β(X). Then there is a neighborhood U of (α,β) such that for all (α,β)U and all nonsplit extensions 0FEG0the object E is λα,β,s-stable if and only if λα,β,s(F)<λα,β,s(G).

This result is stated and proved (for arbitrary Bridgeland stability conditions) in Schmidt [Citation20, Lemma 3.11], and credited there also to Bayer–Macrì [Citation3, Lemma 5.9].

3 Wall and chamber structure

The starting point for the entire discussion that follows is a simple minded observation. Namely, let VP3 be a plane and let Y be the union of a conic in V and a point P also in V. Then there is a short exact sequence (3.1) 0OP3(1)IYIP/V(2)0(3.1)

(read OP3(1) as the ideal of V and IP/V(2) as the relative ideal of YV). If we instead let Y be the union of a conic in V and a point P outside of V then there is a short exact sequence (3.2) 0IP(1)IYOV(2)0(3.2)

(read IP(1) as the ideal of {P}V and OV(2) as the relative ideal of a conic in V). The claim is that in a certain region of the stability manifold of P3, there are exactly two walls with respect to the Chern character ch(IY)=(1,0,2,2), and they are defined precisely by the two pairs of sub and quotient objects appearing in the short exact sequences (3.1) and (3.2).

Mimicking Schmidt’s work for twisted cubics (and their deformations), we argue via tilt stability. Since να,β-stability only involves Chern classes of codimension at least one, and the above two short exact sequences are indistinguishable in codimension one, they give rise to one and the same wall in the tilt stability parameter space. Making this precise is the content of Section 3.1. Moving on to λα,β,s-stability, we apply Schmidt’s method to see that the single να,β-wall “sprouts” two distinct λα,β,s-walls corresponding to (3.1) and (3.2). This is carried out in Section 3.2.

3.1 ν-stability wall

Throughout we specialize to X=P3 with H a (hyper)plane. Let v=(1,0,2,2) be the Chern character of ideal sheaves IY of subschemes YHilb2m+2(P3).

For (α,β)R>0×R we have the tilted abelian category Cohβ(P3) and the slope function να,β. We concentrate on the region β<0, in which any ideal IY of a subscheme YP3 of dimension 1 satisfies μβ(IY)=c1(IY)rk(IY)β=β>0.

As IY is μ-stable also μβ(G)>0 for every quotient IYG and so IYTβ. In particular IYCohβ(P3).

We begin by establishing that there is exactly one tilt-stability wall in the region β<0. The result as well as the argument is analogous to the analysis for twisted cubics by Schmidt [Citation20, Theorem 5.3], except that twisted cubics come with a second wall that destabilizes all objects—for our skew lines there is no such final wall.

Proposition 3.1.

There is exactly one tilt-stability wall for objects with Chern character v=(1,0,2,2) in the region β<0: it is the semicircle W:α2+(β+52)2=(32)2.The wall is defined by exactly the unordered pairs of the following two types:

  1. {IP(1),OV(2)}, where VP3 is a plane and PV, and

  2. {OP3(1),IP/V(2)}, where VP3 is a plane and PV.

    Moreover, the four sheaves figuring in the above unordered pairs are να,β-stable objects in Cohβ(P3) for all (α,β) on W.

The wall W and the hyperbola να,β(v)=0, intersecting at (α,β)=(3/2,5/2), are shown in . Note that we visualize the α-axis as the vertical one.

We first prove the final claim in the proposition. Here is a slightly more general statement:

Lemma 3.2.

  1. Let ZP3 be a finite, possibly empty subscheme. Then the ideal sheaf IZ(1) is a να,β-stable object in Cohβ(P3) for all α>0 and β<1.

  2. Let VP3 be a plane and ZV be a finite, possibly empty subscheme. Then the relative ideal sheaf IZ/V(2) is a να,β-stable object in Cohβ(P3) for all (α,β)R>0×R such that α2+(β+52)2>(12)2.

Remark 3.3.

The condition on (α,β) in part (2) is necessary because of a wall for IZ/V(2). For simplicity let Z be empty. There is a short exact sequence of coherent sheaves 0OP3(3)OP3(2)OV(2)0which yields a short exact sequence 0OP3(2)OV(2)OP3(3)[1]0

in Cohβ(P3) when 3<β<2. The condition να,β(OP3(2))<να,β(OV(2)) is exactly the inequality in (2).

Proof of Lemma 3.2.

The sheaf IZ(1) is μ-stable and satisfies μβ(IZ(1))=1β. For all β<1 it is thus an object in Tβ and so also in Cohβ(P3). Since IZ/V(2) is a torsion sheaf it too belongs to Tβ and so to Cohβ(P3), for all β.

We reduce to the situation Z=. First consider IZ(1) and assume β<1. Note that IZ(1) is a subobject of OP3(1) also in Cohβ(P3) since the torsion sheaf OZ belongs to that category and hence 0IZ(1)OP3(1)OZ0is a short exact sequence in Cohβ(P3). Suppose OP3(1) is να,β-stable. Let FIZ(1) be a proper nonzero subobject in Cohβ(P3) with quotient G. View F also as a subobject of OP3(1), with quotient G. Then να,β cannot distinguish between G and G. Thus if OP3(1) is να,β-stable then να,β(F)<να,β(G)=να,β(G) and so IZ(1) is να,β-stable as well. The reduction from IZ/V(2) to OV(2) is completely analogous.

να,β-stability of the line bundle OP3(1) is a consequence of Δ¯H(OP3(1))=0, by [Citation5, Proposition 7.4.1].

The main task is to establish να,β-stability of OV(2) in the region defined in part (2). By point (3) of [Citation20, Theorem 3.3], the ray β=52 intersects all potential semicircular ν-walls for ch(OV(2)) at their top point, meaning they must be centered at (0,52). All such semicircles of radius bigger than 12 will intersect the ray β=2 (as well as β=3). Thus it suffices to prove that OV(2) is να,β-stable for all α>0 and all integers β.

For such (α,β), suppose 0FOV(2)G0is a short exact sequence in Cohβ(P3) with F0. We claim that ch1β(G)=0. This yields the result, since then να,β(G)= and so OV(2) is να,β-stable.

Let rF=H3ch0(F) and cF=H2ch1(F), i.e. the rank and first Chern class considered as integers. Also let rG=H3ch0(G) and cG=H2ch1(G). By the short exact sequence we have rF+rG=0andcF+cG=1.

The induced long exact cohomology sequence of sheaves shows that H1(F)=0, so from the short exact sequence 0H1(F)[1]FH0(F)0we see that FH0(F) is a coherent sheaf in Tβ. The remaining long exact sequence is 0H1(G)FOV(2)H0(G)0

The map into F cannot be an isomorphism, since H1(G) is in Tβ and F is in Tβ and is nonzero by assumption. Therefore the map in the middle is nonzero and so the rightmost sheaf H0(G) is a proper quotient of OV(2) and so is a torsion sheaf supported in dimension 1. Thus only H1(G) contributes to rG and cG.

Suppose rF0. As FTβ and H1(G)Tβ we have {μβ(F)>0μβ(H1(G))0{cFrFβ>0cGrGβ0{cG1rGβ>0cGrGβ0

and since rG=rF is negative we get 0cGβrG<1. Since these are integers we must have cGβrG=0. Thus ch1β(G)=(cGβrG)H=0as claimed.

If on the other hand rF=0 then also H1(G) has rank zero and hence must be zero as there are no torsion sheaves in Tβ. Thus also G=H0(G) is a sheaf, with vanishing rank and first Chern class. Again ch1β(G)=0 as claimed. This completes the proof. □

By explicit computation (see [Citation20, Theorem 3.3]), all numerical tilt walls with respect to v=(1,0,2,2) in the region β<0 are nested semicircles. More precisely, each is centered on the axis α = 0 and has top point on the curve να,β(v)=0, that is the hyperbola (3.3) β2α2=4.(3.3)

In particular every tilt wall must intersect the ray β=2.

We establish in the following lemma that there is at most one tilt stability wall intersecting the ray β=2 for Chern character v and β<0. We also give the possible Chern characters of sub- and quotient objects that define it. This lemma is tightly analogous to Schmidt [Citation20, Lemma 5.5]. We use an asterisk * to denote an unspecified numerical value.

Lemma 3.4.

Let β0=2 and let α>0 be arbitrary. Suppose there is a short exact sequence 0FEG0of να,β0-semistable objects in Cohβ0(P3) with ch(E)=(1,0,2,*) and να,β0(F)=να,β0(G). Then chβ0(F)=(1,1,12,*)andchβ0(G)=(0,1,12,*) or the other way around.

Proof.

Keep β0=2 throughout. We compute chβ0(E)=(1,2,0,*). Let chβ0(F)=(r,c,d,*) with r,cZ and d12Z. Then chβ0(G)=(1r,2c,d,*).

Since the (very weak) stability function Ztilt sends effective classes to the upper half plane H{0} and Ztilt(E)=Ztilt(F)+Ztilt(G) we have 0IZtilt(F)IZtilt(E).

Since IZtilt=Hch1β0 this gives 0c2.

If c = 0 then να,β0(F)= and να,β0(G)<, which is a contradiction. Similarly if c = 2 then να,β0(F)< and να,β0(G)=, again a contradiction. Therefore c = 1.

With c = 1 we compute να,β0(F)=d12α2randνα,β0(G)=d12α2(1r)and so the condition να,β0(F)=να,β0(G) says (3.4) α2=4d2r1(3.4) so this expression must be strictly positive.

Suppose r1 and apply the Bogomolov inequality (Proposition 2.5) to F: 0Δ¯H(F)=12rdd12r

When r1 this gives d12. On the other hand the positivity of (3.4) gives d > 0 and as d is a half integer this leaves only the possibility d=12 and r = 1.

Similarly suppose r0 and apply the Bogomolov inequality to G: 0Δ¯H(G)=1+2(1r)dd12(1r)

When r0 this gives d12. On the other hand the positivity of (3.4) gives d < 0 and as d is a half integer this leaves only the possibility d=12 and r = 0. □

Proof of Proposition 3.1.

Assume there is a tilt stability wall for v=(1,0,2,2), i.e. there is a short exact sequence 0FEG0of να,β-semistable objects in Cohβ(P3) with ch(E)=(1,0,2,2) and να,β(F)=να,β(G). As already pointed out, the same conditions then hold for some (α,β0) with β0=2. Then by Lemma 3.4, up to swapping F and G, we have (3.5) chβ0(F)=(1,1,12,*)(3.5) (3.6) chβ0(G)=(0,1,12,*).(3.6)

Given any pair F,G of such objects, write out the condition να,β(F)=να,β(G) on (α,β) to obtain the equation for the wall in question; this yields the semicircle as claimed. Thus we have proved that there is at most one tilt-wall and found its equation.

A further result of Schmidt [Citation20, Lemma 5.4] (which requires β to be integral, and so applies for β0=2) says that the only να,β0-semistable objects in Cohβ0(P3) with the invariants (3.5) and (3.6) are FIZ(1)GIZ/V(2)for a finite subscheme ZP3, a plane VP3 and a finite subscheme ZV (where Z and Z are allowed to be empty). Let n and n denote the lengths of Z and Z, respectively. Again for β0=2 we compute ch3β0(F)=ch3β0(IZ(1))=16nch3β0(G)=ch3β0(IZ/V(2))=16n

and moreover ch3β0(E)=23. Thus from ch3β0(E)=ch3β0(F)+ch3β0(G) we find n+n=1and so either Z is empty and (HTML translation failed) is a point, or Z is a point and Z is empty. This proves that only the two listed pairs of semistable objects F,G may occur in a short exact sequence defining the wall.

To finish the proof it only remains to show that both pairs of objects listed do in fact realize the wall. By Lemma 3.2, the sheaves IZ(1) and IZ/V(2) are in Cohβ(P3) and are να,β-semistable (in fact να,β-stable) for all (α,β) on the semicircle. Also, the ideal E=IY of any YHilb2m+2(P3) is an object in Cohβ(P3) (when β<0) and since any ideal is μ-stable it is να,β-stable for α0 (by Proposition 2.6). Hence it is να,β-stable outside the semicircle and at least να,β-semistable on the semicircle. Thus, short exact sequences of the types (3.1) and (3.2) define the wall and we are done. □

3.2 λ-stability walls

Next we apply Schmidt’s Theorem 2.8 to the single να,β-wall found in Proposition 3.1; this yields two λα,β,s-walls.

We set up notation first: for (α,β,s)R>0×R×R>0 we have the doubly tilted category Aα,β(P3) and the slope function λα,β,s. Once and for all we fix an arbitrary value s > 0 and view the (α,β)-plane R>0×R as parameterizing both να,β-stability and λα,β,s-stability; as before we restrict to β<0. Walls and chambers are taken with respect to the Chern character v=(1,0,2,2).

Write PvR>0×R for the open subset defined by να,β(v)>0 and β<0; this is the region to the left of the hyperbola (3.3) in . Theorem 2.8 addresses walls in Pv close to the boundary hyperbola.

Proposition 3.5.

There are exactly two λα,β,s-walls with respect to v=(1,0,2,2) in Pv whose closure intersect the hyperbola (3.3). They are defined exactly by the two pairs of objects listed in Proposition 3.1.

This means that the two walls are (3.7) W1={(α,β)Pv|λα,β,s(OP3(1))=λα,β,s(IQ/V(2))}(3.7) and (3.8) W2={(α,β)Pv|λα,β,s(IP(1))=λα,β,s(OV(2))}(3.8) and the pair of objects defining each wall (close to (α0,β0)) is unique.

We refrain from writing out the (quartic) equations defining them. They do depend on s, but independently of s they both intersect the hyperbola (3.3) in (α0,β0)=(32,52) and as we will show in the following proof, W1 has negative slope at (α0,β0) whereas W2 has positive slope there. Thus W1 lies above W2 (α bigger) in the intersection between Pv and a small open neighborhood of (α0,β0).

Proof.

We apply Schmidt’s theorem 2.8. Firstly, when ch(E)=v we have ch1β(E)=v1βv0=β>0 so the theorem applies. The first part of the theorem says that any λ-wall in Pv, having a point (α0,β0) with να0,β0(v)=0 in its closure, must be defined by one of the two pairs (F,G) listed in Proposition 3.1. This leaves W1 and W2 as the only candidates. Moreover the sub- and quotient objects F and G appearing are να0,β0-stable by Lemma 3.2. Thus the second part of the theorem says that conversely, W1 and W2 are indeed λ-walls, provided they contain points (α,β) arbitrarily close to (α0,β0)=(32,52) such that να,β(F)>0 and να,β(G)>0. It remains to check this last condition.

So let (F,G) be one of the pairs (OP3(1),IP/V(2)) or (IP(1),OV(2)). The region Pv is bounded by the hyperbola να,β(v)=0 and implicit differentiation readily shows that this has slope dαdβ=5/3 at (α0,β0). Similarly να,β(F)=0 has slope –1 and να,β(G)=0 is just the line β=5/2, and in each case να,β>0 is the region to the left of these boundary curves. Thus it suffices to show that our walls have slope >1 at (α0,β0). Now each wall Wi is defined by the condition λα,β,s(F)=λα,β,s(G), which is equivalent to (3.9) (RZ(F))(IZ(G))=(RZ(G))(IZ(F))(3.9) where Z=Zα,β,s is the stability function defined in (2.4) from Section 2.3. Implicit differentiation of this equation at (α0,β0) gives, after some work, that W1 has slope (3.10) (27s16+1)1(1,0)(3.10) and W2 has slope (3.11) (27s4+1)1(0,1)(3.11) both of which are >1, and we are done. □

We are now in position to prove Theorem 1.1. By Proposition 3.5 there exists an open (connected) neighborhood NR>0×R around the β<0 branch of the hyperbola (3.3), such that the only λ-walls in NPv are W1 and W2, defined in (3.7) and (3.8). Moreover it follows from the slopes (3.10) and (3.11) that, after shrinking N further if necessary, W1 lies above (α bigger) W2 throughout NPv. Thus the two walls separate N into three chambers, which we label I, II, and III in order of decreasing α.

Proof of Theorem 1.1

(I). By Proposition 3.1 there is a single semicircular wall (in the region β<0) for ν-stability. It follows from Theorem 2.7 that the class of λα,β,s-stable objects in Aα,β(P3) for (α,β) in chamber I coincides with the class of να,β-stable objects in Cohβ(P3) for (α,β) outside the single ν-wall (up to shrinking N even further if necessary).

Moreover, for α sufficiently big, the ν-stable objects in Cohβ(P3) are exactly the μ-stable coherent sheaves (Proposition 2.6). For Chern character v=(1,0,2,2) these are the ideals IY with YHilb2m+2(P3). □

Proof of Theorem 1.1

(II). Let E be λα,β,s-stable for (α,β) in chamber II. Since semistability is a closed property, E is semistable on the wall W1. If E is stable on the wall, then it is also stable in chamber I hence it is an ideal sheaf in Hilb2m+2(P3) by part (I). Such an ideal remains stable on the wall if and only if it is not an extension of the type (3.1), that is if and only if it is the ideal of a nonplanar subscheme. This is case (II)(i) in the Theorem.

If on the other hand E is stable in chamber II, but strictly semistable on W1, then by Proposition 3.5 it is a nonsplit extension of the pair (3.12) {OP3(1),IP/V(2)}(3.12) and we determine the direction of the extension (which object is the subobject and which is the quotient) as follows: we claim that (3.13) λα,β,s(IP/V(2))<λα,β,s(OP3(1))(3.13) for all (α,β) in chamber II sufficiently close to (α0,β0). Granted this, it follows that for E to be stable in chamber II it must be a nonsplit extension as in case (II)(ii) in the Theorem. Conversely it follows from Proposition 2.9 that every such nonsplit extension is indeed stable in chamber II. To verify (3.13) we let (3.14) Φ(α,β)=RZ(F)IZ(G)RZ(G)IZ(F)(3.14) with Z=Zα,β,sF=OP3(1) and G=IP/V(2). Thus W1 is defined by Φ(α,β)=0 and (3.13) is equivalent to Φ(α,β)<0. It thus suffices to check that the partial derivative of Φ with respect to α is positive at (α0,β0). An explicit computation yields in fact Φα(α0,β0)=2+27s8>0.

It remains only to show uniqueness of the nonsplit extensions FP,V, that is dimExtP31(OP3(1),IP/V(2))=1.

But this space is H1(IP/V(1)), which is isomorphic to H0(k(P))=k via the short exact sequence 0IP/V(1)OV(1)k(P)0.

Proof of Theorem 1.1

(III). Let E be λα,β,s-stable for (α,β) in chamber III. Since semistability is a closed property, E is semistable on the wall W2. If E is stable on W2, then it is stable in chamber II. This means two things: first, by part (II) of the Theorem E is either an ideal sheaf of a nonplanar subscheme or a nonsplit extension FP,V as in case (II)(ii). Second, to remain stable on W2 the object E cannot be in a short exact sequence of the type (3.2) ruling out ideal sheaves of plane conics union a point. Also the sheaves FP,V sit in short exact sequences of this type, as we show in Lemma 3.6 (the vertical short exact sequence in the middle), and so are ruled out as well. Hence E is an ideal sheaf of a disjoint pair of lines as claimed in (III)(i).

If on the other hand E is strictly semistable on W2, then by Proposition 3.5 E is a nonsplit extension (in either direction) of the pair {IP(1),OV(2)}.

Now we claim that λα,β,s(OV(2)<λα,β,s(IP(1))for all (α,β) in chamber III sufficiently close to (α0,β0). We prove this as in part II above, by partial differentiation of Φ defined in (3.14), this time with F=IP(1) and G=OV(2). We find Φα(α0,β0)=12+27s8>0.

As before we conclude that E is a nonsplit extension as in (III)(ii) and by Proposition 2.9 all such extensions are stable.

It remains to verify uniqueness of the extensions GP,V, i.e. dimExtP31(IP(1),OV(2))=1when PV. For this first apply Hom(,OV(1)) to the short exact sequence 0IPOP3k(P)0

to obtain a long exact sequence which together with the vanishing of H1(OV(1)) and H2(OV(1)) gives an isomorphism ExtP31(IP,OV(1))Ext2(k(P),OV(1))and ignoring twists, as these are not seen by k(P), the right hand side is Serre dual to Ext1(OV,k(P)). This is one dimensional as is seen by applying Hom(,k(P)) to the sequence 0OP3(1)OP3OV0.

3.3 The special sheaves

Let F=FP,V and G=GP,V denote sheaves given by nonsplit extensions of the form (1.1) and (1.2), respectively. The definition through (unique) nonsplit extensions is indirect and it is useful to have alternative constructions available. We give such constructions here and compute the spaces of first order infinitesimal deformations.

Lemma 3.6.

There is a commutative diagram with exact rows and columns as follows:

[[INLINE FIGURE]]

Proof.

Up to identifying the skyscraper sheaf k(P) with any of its twists, there are canonical short exact sequences as in the bottom row and the rightmost column. The diagram can then be completed by letting F be the fiber product as laid out by the square in the bottom right corner. It remains only to verify that the middle row is nonsplit. But if it were split the middle column twisted by OP3(1) would be a short exact sequence of the form 0IPIP/V(1)OP3OV(1)0.

Taking global sections this yields a contradictory left exact sequence in which all terms vanish except H0(OP3)=k. □

Proposition 3.7.

We have dimExt1(F,F)=11.

Proof.

We will actually only prove that the dimension is at most 11. The opposite inequality may be shown by similar techniques, although it follows from viewing Ext1(F,F) as a Zariski tangent space to the 11-dimensional moduli space MII studied in the next section.

Apply Hom(,F) to the middle row in the diagram in Lemma 3.6. This yields a long exact sequence H1(F(1))Ext1(F,F)Ext1(IP/V(2),F)H2(F(1))and from the middle column of the diagram we compute H1(F(1))=H2(F(1))=0. Thus we proceed to show that dimExt1(IP/V(2),F)11.

Apply Hom(IP/V(2),) to the middle row in the diagram. This yields a long exact sequence: Ext1(IP/V,IP/V)Ext1(IP/V(2),F)Ext1(IP/V(1),OP3)

The space on the right is Serre dual to H2(IP/V(5))H2(OV(5)), which again on V is Serre dual to H0(OV(2)). This has dimension 6. At least heuristically, the space on the left should have dimension 5, as it may be viewed as a tangent space to the incidence variety IP3×Pˇ3 seen as a moduli space for the sheaves IP/V. More directly we may apply Hom(IP/V,) to the Koszul complex on V 0OV(2)OV(1)2IP/V0to obtain a long exact sequence Ext1(IP/V,OV(1))22·2Ext1(IP/V,IP/V)Ext2(IP/V,OV(2))1 where the indicated dimensions may be computed by applying Hom(IP/V(d),) (for d = 1, 2) to the sequence 0OP3(1)OP3OV0.

We skip further details. It follows then that dimExt1(IP/V,IP/V)5 and so dimExt1(IP/V(2),F) is at most 5+6=11. □

Lemma 3.8.

There is a commutative diagram with exact rows and columns as follows:

[[INLINE FIGURE]]

Proof.

From the Euler sequence (3.15) 0ΩV1OV(1)3OV0(3.15) on VP2 it follows that ΩV1(2) has a unique section (up to scale) vanishing at PV. This leads to the short exact sequence in the bottom row. Moreover there is a canonical short exact sequence as in the rightmost column. The rest of the diagram can then be formed by taking G to be the fiber product as laid out by the bottom right square. It just remains to verify that the middle row is indeed nonsplit. But if it were split the middle column would be a short exact sequence of the form 0OP3(2)OV(2)IP(1)ΩV10.

This sequence implies that H1(ΩV1(1)) is isomorphic to H1(IP(2)), which is one dimensional. But the Euler sequence shows that in fact H1(ΩV1(1))=0. □

Proposition 3.9.

We have dimExt1(G,G)=8.

Proof.

We will be using the short exact sequence (3.16) 0OP3(2)GΩV10(3.16) which sits as the middle column in Lemma 3.8. As preparation we observe that all (dimensions of) Hi(ΩV1(d)) may be computed from the Euler sequence, and this enables us to compute several Hi(G(d)) from (3.16). We use these results freely below without writing out further details.

Apply Hom(,G) to (3.16) to produce a long exact sequence 0Hom(ΩV1,G)0Hom(G,G)1H0(G(2))4Ext1(ΩV1,G)Ext1(G,G)H1(G(2))0in which we have indicated some of the dimensions: Hi(G(2)) are computed from (3.16) as sketched above, and since G is simple we have Hom(G,G)=k. For the same reason (3.16) is nonsplit, which implies Hom(ΩV1,G)=0. It thus remains to see that the dimension of Ext1(ΩV1,G) is 11.

Next apply Hom(,G) to the Euler sequence. This gives a long exact sequence Hom(ΩV1,G)0Ext1(OV,G)1Ext1(OV,G(1))34·3Ext1(ΩV1,G)Ext2(OV,G)0where again dimensions have been indicated: the vanishing of Hom(ΩV1,G) has already been noted, and there remain several spaces of the form Exti(OV,G(d)). These may be computed from Hi(G(d)) and the long exact sequence resulting from applying Hom(,G(d)) to 0OP3(1)OP3OV0.

It follows that dimExt1(ΩV1,G)=11 and we are done. □

4 Moduli spaces and universal families

By the classification of stable objects in chamber II, the moduli space MII is at least obtained as a set from Hilb2m+2(P3)=CS by just replacing the divisor EC, parameterizing conics union a point inside a plane, with the incidence variety I, parameterizing just pairs (P, V) of a point P inside a plane V. Similarly, the moduli space MIII of stable objects in chamber III is obtained from MII set-theoretically by removing MIIS and replacing the divisor FS, parameterizing pairs of incident lines with a spatial embedded point at the intersection, with the incidence variety I. We shall carry out each of these replacements as a contraction, i.e. a blow-down, and prove that this indeed yields MII and MIII, essentially by writing down a universal family for each case.

4.1 The contraction CC

Recall that C is isomorphic to the blow-up of P3×Hilb2m+1(P3) along the universal curve Z, where Hilb2m+1(P3) is the Hilbert scheme of plane conics in P3 [Citation13]. The exceptional divisor E is comprised of plane conics with an embedded point.

It is helpful to keep an eye at the following diagram

[[INLINE FIGURE]] (4.1)where π sends a conic CHilb2m+1(P3) to the plane VPˇ3 it spans and b is the blowup along the universal family Z of conics.

Remark 4.1.

It will sometimes be useful to resort to explicit computation in local coordinates. For this let UPˇ3 be the affine open subset of planes VP3 with equation of the form (4.2) x3=c0x0+c1x1+c2x2.(4.2)

Furthermore the P5 of symmetric 3 × 3 matrices (sij) parameterizes plane conics (4.3) 0i,j2sijxixj=0(4.3) so that Hilb2m+1(P3)|UP5×U with universal family defined by the two Equationequations (4.2) and Equation(4.3). This is also the center for the blowup b, and we note that it is nonsingular.

Lemma 4.2.

π is a Zariski locally trivial P5-bundle. More precisely, let IP3×Pˇ3 be the incidence variety and let E=pr2*(pr1*OP3(2)|I).

Then E is locally free of rank 6 and Hilb2m+1(P3)P(E) over Pˇ3.

Here P(E) denotes the projective bundle parameterizing lines in the fibers of E. Starting with the observation that the fiber of E over VPˇ3 is H0(V,OV(2)) (note that H1(V,OV(2))=0, so base change in cohomology applies) the Lemma is straight forward and we refrain from writing out details.

Now let EC be the locus of planar YC. The condition on a disjoint union Y=C{P} to be in E is just that P is in the plane V spanned by C. For a conic with an embedded point the condition YV also singles out the scheme structure at the embedded point. View E as a variety over the incidence variety IP3×Pˇ3 via the morphism (idP3×π)°b (refer to for simple illustrations of the types of elements in E and E. In fact, we find it helpful for keeping track of the various divisors introduced in this section.)

Fig. 3 A Circle represents a conic contained in a plane that is shown as a parallelogram, and a red dot is a point, possibly embedded in the conic. The arrow is the direction vector at an embedded point. Note that in the left illustration, the arrow is strictly contained in the plane.

Fig. 3 A Circle represents a conic contained in a plane that is shown as a parallelogram, and a red dot is a point, possibly embedded in the conic. The arrow is the direction vector at an embedded point. Note that in the left illustration, the arrow is strictly contained in the plane.

Proposition 4.3.

E is a Zariski locally trivial P5-bundle over the incidence variety IP3×Pˇ3. The restriction OC(E)|P5 to a fiber is isomorphic to OP5(1).

Before giving the proof, we harvest our application:

Corollary 4.4.

There exist a smooth algebraic space C, a morphism ϕ:CC and a closed embedding IC, such that ϕ restricts to an isomorphism from CE to CI and to the given projective bundle structure EI. Moreover ϕ is the blowup of C along I.

It is well known that the condition verified in Proposition 4.3 implies the contractibility of E/I in the above sense. In the category of analytic spaces this is the Moishezon [Citation16] or Fujiki–Nakano [Citation10, Citation17] criterion. In the category of algebraic spaces the contractibility is due to Artin [Citation2, Corollary 6.11], although the statement there lacks the identification with a blowup. Lascu [Citation12, Théorème 1] however shows that once the contracted space C as well as the image IC of the contracted locus are both smooth, it does follow that the contracting morphism is a blowup. Strictly speaking Lascu works in the category of varieties, but our C turns out to be a variety anyway:

Remark 4.5.

The algebraic space C is in fact a projective variety. We prove this in Section 5 using Mori theory. As the arguments there and in the present section are largely independent we separate the statements.

We also point out that the smooth contracted space C is unique once it exists: in general, suppose ϕ:XU and ψ:XV are proper birational morphisms between irreducible separated algebraic spaces (say, of finite type over k) with U and V normal. Moreover assume that ϕ(x1)=ϕ(x2) if and only if ψ(x1)=ψ(x2). Let Γ denote the image of (ϕ,ψ):XU×V. Then each of the projections from Γ to U and V is birational and bijective and hence an isomorphism by Zariski’s Main Theorem (for this in the language of algebraic spaces we refer to the Stacks Project [Citation21, Tag05W7]).

Proof of Proposition 4.3.

Consider the divisor E¯={(P,C)P3×Hilb2m+1(P3)|P in the plane spanned by C}.

It follows from Lemma 4.2 that P3×Hilb2m+1(P3)idP3×πP3×Pˇ3is a P5-bundle, hence its restriction to (idP3×π)1(I)=E¯ is a P5-bundle over IP3×Pˇ3.

Now, E is the strict transform of E¯, i.e. its blow-up along ZE¯. But this is a Cartier divisor, since E¯ is smooth, and so EE¯. This proves the first claim.

Again using that E¯ is smooth, its strict transform E satisfies the linear equivalence (4.4) E=b*(E¯)E.(4.4)

The term b*(E¯)=b*(idP3×π)*(I) is a pullback from the base of the P5-bundle, so its restriction to any fiber is trivial. Thus it suffices to see that E restricted to a fiber P5 is a hyperplane. Now the isomorphism b:EE¯ identifies EEE with ZE. In the local coordinates from Remark 4.1 the divisor E is given by Equationequation (4.2) and Z is given by the additional Equationequation (4.3). Here (sij) are the coordinates on the fiber P5 and clearly (4.3) defines a hyperplane in each fiber—it is the linear condition on the space of plane conics given by passage through a given point. □

Remark 4.6.

The locus CS consists of pairs of intersecting lines with a spatial embedded point at the intersection (and, as degenerate cases, planar double lines with a spatial embedded point). On the other hand E consists only of planar objects, so E is disjoint from S. Thus we may extend the contraction ϕ to a morphism between algebraic spaces (ϕid):CSCS

which is an isomorphism away from E and restricts to the P5-bundle EI as before.

4.2 Moduli in chamber II

In this section we shall modify the universal family on the Hilbert scheme Hilb2m+2(P3)=CS in such a way that we replace its fibers over EC with the objects FP,V in Theorem 1.1. This family induces a morphism CSMIIand we conclude via uniqueness of normal (in this case smooth) contractions that MII coincides with CS.

Here is the construction: let YP3×Hilb2m+2(P3)

be the universal family and let VP3×Ebe the E-flat family whose fiber VξP3 over a point ξ mapping to (P,V)I is the plane V. Clearly V can be written down as a pullback of the universal plane over Pˇ3. We view V as a closed subscheme of P3×Hilb2m+2(P3). Then our modified universal family is the ideal sheaf IYV.

Remark 4.7.

We emphasize the (to us, at least) unusual situation that IYV is the ideal of a very nonflat subscheme, yet as we show below it is flat as a coherent sheaf. Its fibers over points in E are not ideals at all, but rather the objects FP,V.

Theorem 4.8.

As above let Y be the universal family over Hilb2m+2(P3) and V the family of planes in P3 parameterized by E.

  1. IYV is flat as a coherent sheaf over Hilb2m+2(P3). Its fibers IYVk(ξ) over ξHilb2m+2(P3) are stable objects for stability conditions in chamber II.

  2. The morphism Hilb2m+2(P3)MII determined by IYV induces an isomorphism CSMII.

We begin by showing that the fibers IYVk(ξ) over ξE sit in a short exact sequence of the type (1.1). The mechanism producing such a short exact sequence is quite general. Note that when Y=Yξ is a conic with a (possibly embedded) point P in a plane V=Vξ, we have IY/VIP/V(2) and IVOP3(1).

Lemma 4.9.

Let X be a projective scheme, YX×S an S-flat subscheme, ES a Cartier divisor and VE×S an E-flat subscheme such that Y(E×S)V. Let ξE. Then there is a short exact sequence 0IYξ/VξIYVk(ξ)IVξ0.

In particular if S is integral in a neighborhood of E then IYV is flat over S.

Proof.

Observe that the last claim is implied by the first: outside of E the ideal IYV agrees with IY, which is flat. For ξE we have YξVξ and so a short exact sequence 0IVξIYξIYξ/Vξ0.

The short exact sequence in the statement has the same sub and quotient objects in opposite roles, so the Hilbert polynomial of IYVk(ξ) agrees with that of IYξ. Thus the Hilbert polynomial of the fibers of IYV is constant; over an integral base this implies flatness.

Begin with the short exact sequence 0IYVOX×SOYV0

and tensor with OX×E to obtain the exact sequence (4.5) 0Tor1X×S(OYV,OX×E)IYV|EOX×EOV0.(4.5)

The kernel of the rightmost map is clearly the ideal IVOX×E. To compute the Tor-sheaf on the left use the short exact sequence 0OS(E)OSOE0.

Pull this back to X × S and tensor with OYV to see that Tor1X×S(OYV,OX×E) is isomorphic to the kernel of the homomorphism OYV(pr2*E)OYV

which locally is multiplication by an equation for E. Thus Tor1X×S(OYV,OX×E)J(pr2*E)where JOYV is the ideal locally consisting of elements annihilated by an equation for E.

We compute J in an open affine subset SpecAX×S in which Y and V are given by ideals IY and IV respectively and fA is (the pullback of) a local equation for E. Thus J corresponds to (IYIV:f)/(IYIV). Now fIV since VX×E. This implies that for gA the condition fgIY is equivalent to fgIYIV and so (IYIV:f)=(IY:f). Moreover the latter equals IY, since Y is flat over S, so that multiplication by the non-zero-divisor f remains injective after tensor product with OY, that is A/IY. Thus J is locally (IYIV:f)/(IYIV)=IY/(IYIV)(IY+IV)/IV=IYE/IVwhere we write YE for the restriction Y(X×E)=YV. This shows Tor1X×S(OYV,OX×E)IYE/V(pr2*E)

and (4.5) gives the short exact sequence 0IYE/V(pr2*E)IYV|EIV0on X × E. Finally restrict to the fiber over a point ξE: since YE and V are both E-flat this yields the short exact sequence in the statement. □

Lemma 4.9 does not guarantee that the short exact sequence obtained is nonsplit. Showing this in the case at hand requires some work. Our strategy is to exhibit a certain quotient sheaf IYVk(ξ)Q and check that the split extension IVξIYξ/Vξ admits no surjection onto Q. In fact Q=OV(2) will work:

Lemma 4.10.

Let VP3 be a plane and PV a point. Then there is no surjection from OP3(1)IP/V(2) onto OV(2).

Proof.

Just note that Hom(IP/V(2),OV(2))=k is generated by the (nonsurjective) inclusion, whereas Hom(OP3(1),OV(2))=0. □

We will produce the required quotient sheaf by the following construction, which depends on the choice of a tangent direction at ξ in S:

Lemma 4.11.

With notation as in Lemma 4.9, let T=Speck[t]/(t2) and let TS be a closed embedding such that TE is the reduced point {ξ}. Let Y=Yξ and V=Vξ.

Define a subscheme YY by the ideal (I:t)/(t)OVwhere IOV×T is the ideal of Y(V×T). Then there is a surjection IYVk(ξ)IY/V.

Remark 4.12.

Since t2=0 we trivially have t(I:t). Since also I(I:t) we furthermore have YY. If we extend T to an actual one parameter family of objects Yt, we may think of Y as the limit of YtV as t0, in other words it is the part of Y that remains in V as we deform along our chosen direction.

Proof.

Let YTP3×T denote the restriction of Y to T. We claim that IY/V is isomorphic to the relative ideal of YT(V×{ξ}) in YT(V×T). Assuming this, there are surjections IYVOTIYT(V×{ξ})IY/V

(the middle term is the ideal of (YV)|T=YT(V×{ξ}) as a subscheme of P3×T). Restriction to the fiber over ξ gives the surjection in the statement.

To prove the claim, we first observe that for any two subschemes A and B of some ambient scheme, there is an isomorphism I(AB)/AIB/(AB)between the relative ideal sheaves; this is the identity (I+J)/IJ/(IJ) between quotients of ideals. Apply this to A=V×T,B=YT(V×{ξ})

so that AB=YT(V×T)AB=(YT(V×{ξ}))(V×T)=(Y(V×T))(V×{ξ}).

The claim as stated thus says that IY/V is isomorphic to IB/AB, and we are free to replace the latter by I(AB)/A.

Next let SpecR be an affine open subset in V and IR[t]/(t2) the ideal defining Y(V×T) there. Locally the ideal I(AB)/A is then I(t)R[t]/(t2). Now multiplication with t is an isomorphism of R[t]/(t2)-modules (I:t)/(t)I(t).

The left hand side is precisely IY/V in the open subset SpecR. □

Proof of Theorem 4.8

(i). By Lemma 4.9 it suffices to show that the fiber of IYVk(ξ) over ξE is isomorphic to FP,V. Moreover, for such ξ, the same Lemma yields a short exact sequence 0IP/V(2)IYVk(ξ)OP3(1)0and since FP,V is the unique such nonsplit extension it is enough to show that the above extension is nonsplit. In view of Lemma 4.10 this follows once we can show the existence of a surjection IZWk(ξ)OV(2).

For this it suffices, in the notation of Lemma 4.11, to choose THilb2m+2(P3) such that the subscheme YV is a conic.

Nondegenerate case. First assume Y=Yξ is a disjoint union Y=C{P} of a conic CV and a point PV. Consider the one parameter family Yt=C{Pt} in which the conic part C is fixed while the point Pt travels along a line intersecting V in the point P. In suitable affine coordinates we may take V to be the plane z = 0 in A3=Speck[x,y,z], the point P to be the origin and C to be given by some quadric q=q(x,y) not vanishing at P. Let the one parameter family over Speck[t] consist of the union of C with the point Pt=(0,0,t). This is given by the ideal (q,z)(x,y,zt)=(xq,yq,(zt)q,xz,yz,(zt)z).

Now restrict to T=Speck[t]/(t2) and intersect the family with V × T. The resulting subscheme is defined by the ideal I=(xq,yq,(zt)q,xz,yz,(zt)z)+(z)=(xq,yq,tq,z)and (I:t)/(t)=(q,z) which defines CV. Thus Y=C and we are done.

Embedded point with nonsingular support. Suppose Y is a conic CV with an embedded point supported at a point P in which C is nonsingular, where the normal direction corresponding to the embedded point is along V. Then take the one parameter family in which C and the supporting point P is fixed and the embedded structure varies in the P1 of normal directions. In suitable affine coordinates we may take V to be the plane z = 0 in A3=Spec k[x,y,z], P to be the origin and C given by a quadric q=q(x,y) vanishing at P and with, say, linear term y. Take the one parameter family of C with an embedded point given by (q,z)(x,y2,zty)=(xq,yq,zq,ztq)

(the equality requires some computation). After intersection with V × T this gives I=(xq,yq,tq,z)and (I:t)/(t)=(q,z). This is C.

Embedded point at a singularity. Let CV be the union of two distinct lines intersecting in P and consider a planar embedded point at P. Despite the singularity, there is still a P1 of embedded points at P. We take this to be our one parameter family, i.e. we deform the embedded point structure away from the planar one.

In local coordinates we take P to be the origin in A3 and C to be the union of the x- and y-axes in the xy-plane V. Then (xy,z)(x,y,z)+(ztxy)=(xy2,x2y,ztxy)is our one parameter family of embedded points at the origin, with t = 0 corresponding to the planar embedded point. The intersection with V × T is given by I=(xy2,x2y,z,txy) and (I:t)/(t)=(xy,z). This is C.

Embedded point in a double line. Let CV be a planar double line together with a planar embedded point at PC and take the one parameter family of embedded points in P.

In local coordinates we take P to be the origin in A3 and C to be V(z,y2). Then (z,y2)(x,y,z)+(zty2)=(y3,xy2,zty2)is our one parameter family of embedded points at the origin, with t = 0 corresponding to the planar embedded point. The intersection with V × T is given by I=(xy2,y3,z,ty2)

and (I:t)/(t)=(z,y2). This is C. □

Proof of Theorem 4.8

(ii). The morphism Hilb2m+2(P3)MII is clearly an isomorphism away from E, and it sends ξE (lying over (P,V)I) to FP,V, which determines and is uniquely determined by (P, V). Moreover MII is smooth at these points by Proposition 3.7. The claim follows from uniqueness of normal contractions. □

4.3 Moduli in chamber III

In this section we show that the moduli space MIII is a contraction of S. The argument parallels that for MII closely.

Let FS be as in Notation 1.2. Thus an element YS is either a pair of intersecting lines with a spatial embedded point at the intersection, or as degenerate cases, a planar double line with a spatial embedded point. It is in a natural way a P2-bundle over the incidence variety IP3×Pˇ3 via the map FIthat sends Y to the pair (P, V) consisting of the support PY of the embedded point and the plane V containing Y{P}. In parallel with Proposition 4.3 one may show that OS(F) restricts to OP2(1) in the fibers of F/I and so there is a contraction (4.6) ψ:SS(4.6) to a smooth algebraic space S, such that ψ is an isomorphism away from F and restricts to the P2-bundle FI. However, in this case we can be much more concrete thanks to the work of Chen–Coskun–Nollet [Citation8], where birational models for S are studied in detail (and in greater generality: moduli spaces for pairs of codimension two linear subspaces of projective spaces in arbitrary dimension). The following proposition is [Citation8, Theorem 1.6 (4)]; we sketch a simple and slightly different argument here.

Proposition 4.13.

There is a contraction as in (4.6) where S is the Grassmannian G(2, 6) of lines in P5.

Proof.

First consider an arbitrary quadric QPn. Any finite subscheme in Q of length 2, reduced or not, determines a line in Pn. This defines a morphism (4.7) Hilb2(Q)G(2,n+1).(4.7)

It is clearly an isomorphism away from the locus in G(2,n+1) consisting of lines contained in Q. On the other hand, over every element of G(2,n+1) defining a line contained in Q, the fiber is the P2 consisting of length two subschemes of that line.

Apply the above observation to the (Plücker) quadric Q=G(2,4) in P5, so that SHilb2(Q) (see Section 2.1). For every plane VP3 and every point PV, the pencil of lines in V through P defines a line in Q=G(2,4) and in fact every line is of this form. The fiber of (4.7) above such an element of G(2, 5) consists of all pairs of lines in V intersecting at P. It follows that (4.7) is the required contraction SS. □

Remark 4.14.

Chen–Coskun–Nollet furthermore shows that (4.6) is a K-negative extremal contraction in the sense of Mori theory. In fact, S is Fano and its Mori cone is spanned by two rays. Either ray is thus contractible; one contraction is (4.6) and the other is the natural map to the symmetric square of the Grassmannian of lines in P3. This statement is extracted from Theorem 1.3, Lemma 3.2, and Proposition 3.3 in loc. cit. Inspired by this work we return to the Mori cone of the conics-with-a-point component C in Section 5.

We proceed as for chamber II by modifying the universal family of pairs of lines in order to identify the moduli space MIII with the contracted space S. Let YP3×Sbe the restriction of the universal family over Hilb2m+2(P3) to the component S. Moreover, there is a flat family over the incidence variety IP3×Pˇ3 whose fiber over (P, V) is the plane V with an embedded point at P. Pull this back to F to define a family WP3×F.

We argue as in Section 4.2 but with the family of planes V replaced by the family of planes with an embedded point W.

Theorem 4.15.

Let Y and W be as above.

  1. IYW is flat as a coherent sheaf over S. Its fibers IYWk(ξ) over ξS are stable objects for stability conditions in chamber III.

  2. The morphism SMIII determined by IYW induces an isomorphism SMIII.

For ξF lying over (P, V) we have IYξ/WξOV(2) and IWξIP(1). Thus Lemma 4.9 yields a short exact sequence 0OV(2)IYWk(ξ)IP(1)0and we show that it is nonsplit by exhibiting a certain quotient sheaf of IYWk(ξ). This time we use IQ/V(1) where QV is a point distinct from P.

Lemma 4.16.

Let VP3 be a plane and P,QV two distinct points. There is no surjection from OV(2)IP(1) to IQ/V(1).

Proof.

Every nonzero homomorphism OV(2)OV(1)

has image of the form IL/V(1) where LV is a line, whereas every nonzero homomorphism IP(1)OV(1)has image IP/V(1). Thus any nonzero homomorphism from the direct sum of these two sheaves has image I(1) where IOV is one of IL/VIP/V or their sum IL/V+IP/V={IP/Vif PLOVotherwise.

Thus the image is never IQ/V(1) for QP. □

Proof of Theorem 4.15.

The proof for Theorem 4.8 carries over; we only need to detail the construction of quotient sheaves via one parameter families. As before we write down families over A1=Speck[t] and then restrict to T=Speck[t]/(t2). We then apply Lemma 4.11, with W in the role of the family denoted V in the Lemma. The outcome of Lemma 4.11 will be a quotient sheaf of the form IY/W, where W=Wξ is a plane V with an embedded point at P. We end by intersecting with V to produce a further quotient of the form IYV/V. We shall choose one parameter families such that the latter is isomorphic to IQ/V(1) with QP.

Distinct lines. Let C=LL0 be a pair of distinct lines inside V intersecting at P. Choose another plane V containing L0 and a point QL0 distinct from P. The pencil of lines LtV through Q yields a one parameter family Zt=LLtof disjoint pairs of lines for t0, with flat limit Z0W being C with a spatial embedded point at P.

In suitable affine coordinates A3 let V be V(z), let P be the origin and let C=V(z,xy). Then W=V(xz,yz,z2). Furthermore let Q=(0,1,0) and Lt=V(x,zt(y1)). This leads to the family Z defined by the ideal (y,z)(x,zt(y1))=(xy,xz,(zt(y1))y,(zt(y1))z)and the intersection with W × T is given by I=(xz,yz,z2,xy,ty(y1),tz)

Thus (I:t)/(t)=(z,xy,y(y1)), which defines the union of the x-axis and the point Q. This is YW and thus IYV/V=IY/V is isomorphic to IQ/V(1).

Double lines. Let CV be a double line inside the plane V with PC. We shall define an explicit one parameter family with central fiber Y0W being C with a spatial embedded point at P.

Geometrically, the family is this: let LV be the supporting line of C. Consider a line MV not through P and let Q be its intersection point with L. Also let M be a line through P and not contained in V. Let Rt be a point on M moving toward Q as t0, and let Rt be a point on M moving toward P, but much faster than Rt moves (quadratic versus linearly). Then let Lt be the line through Rt and Rt and let Yt=LLt for t0.

Let P be the origin in suitable affine coordinates A3, let V be the xy-plane V(z) and let CV be the double x-axis V(y2,z). Thus Y0 corresponds to (z,y2)(x,y,z)2=(xz,yz,z2,y2).

Now let L=V(y,z), let Lt be the line through (1,t,0) and (0,0,t2), that is Lt=V(txy,ty+zt2)and take Yt=LLt for t0. This yields the family (the following identity requires a bit of fiddling) (y,z)(txy,ty+zt2)=((txy)y,(txy)z,(ty+zt2)z,xz+ty(x1)).

Reducing this modulo t gives the original Y0. The intersection with W × T gives I=(xz,yz,z2,(txy)y,ty(x1))and so YW is defined by (I:t)/(t)=(xz,yz,z2,y2,y(x1))=(y,z)(x1,y2,z)(x,y,z2).

This is the line L with an embedded point at Q (inside V) and another embedded point along the z-axis at P. Intersecting with V removes the embedded point at P, leaving the line L with an embedded point at Q. Thus IYV/VIQ/V(1).

This establishes part (i) precisely as in the proof of Theorem 4.8 and part (ii) then follows by smoothness of MIII (from Proposition 3.9) and by uniqueness of normal (here smooth) contractions. □

5 The Mori cone of C and extremal contractions

In this final section we shall prove that CC is the contraction of a K-negative extremal ray in the Mori cone. It follows that the contracted space C is projective.

To set the stage we recall the basic mechanism of K-negative extremal contractions. Let X be a projective normal variety and α a curve class (modulo numerical equivalence) which spans an extremal ray in the Mori cone. If also the ray is K-negative, i.e. the intersection number between α and the canonical divisor KX is negative, then there exists a unique projective normal variety Y and a birational morphism f:XY which contracts precisely the effective curves in the class α.

5.1 Statement

We denote elements in C by the letter Y. It is the union of a (possibly degenerate) conic denoted C and a point denoted P. If the point is embedded, PC denotes its support. We also write VP3 for the unique plane containing C.

Define four effective curve classes (modulo numerical equivalence) on C. Each is described as a family {Yt}, and we use a subscript t to indicate a parameter on the piece that varies (all choices are to be made general, e.g. C nonsingular unless stated otherwise, etc.):

δ:fix a conic C and a point PC. Let Yt be C with an embedded point at P, varying in the P1 of normal directions to CP3 at P.

ϵ:fix a plane V, a conic CV and a line LV. Let Pt vary along L and let Yt=C{Pt}.

ζ:fix a plane V, a pencil of conics CtV and a point PV. Let Yt=Ct{P}.

η:fix a line L and a point PL. Let Vt be the pencil of planes containing L and let Ct be the planar double structure on L inside Vt. Then let Yt be Ct with an embedded spatial point at P.

In ϵ there are implicitly two elements with an embedded point, namely where L intersects C. Similarly there is one element in ζ with an embedded point, corresponding to the pencil member Ct that contains P.

Theorem 5.1.

The Mori cone of C is the cone over a solid tetrahedron, with extremal rays spanned by the four curve classes δ,ϵ,ζ,η. Of these, the first three are K-negative, whereas η is K-positive. The contraction corresponding to ζ is CC.

Corollary 5.2.

C is projective.

Remark 5.3.

The last claim in the theorem is clear: by contracting ζ we forget the conic part of YV, keeping only V and the point PV. By uniqueness of (normal) contractions the contracted variety is C. Also, with reference to Diagram Citation4.Citation1 (from Section 4.1), the contraction of δ is the blowing down b. The theorem furthermore reveals a third K-negative extremal ray spanned by ϵ. The corresponding contraction has the effect of forgetting the point part of YV, keeping only the conic; thus the contracted locus in C is the same as for ζ, but the contraction happens in a “different direction.” We do not know if the contracted space has an interpretation as a moduli space for Bridgeland stable objects.

5.2 The canonical divisor

Use notation as in Diagram Citation4.Citation1 and Lemma 4.2. We read off that the Picard group of C has rank 4 and is generated by the pullbacks of the following divisor classes: HP3 a plane,HPˇ3 a plane in the dual space,A=c1(OP(E)(1)),EC the exceptional divisor for the blowup b.

Moreover numerical and linear equivalence of divisors coincide on C. Here we only use that the Picard group of a projective bundle over some variety X is Pic(X)Z, with the added summand generated by O(1), and the Picard group of a blowup of X is Pic(X)Z, with the added summand generated by the exceptional divisor.

As long as confusion seems unlikely to occur we will continue to use the symbols H, H and A for their pullbacks to C, or to an intermediate variety such as P3×Hilb2m+2(P3) in Diagram Citation4.Citation1.

Lemma 5.4.

The canonical divisor class of C is KC=4H8H6A+E.

Proof.

This is a standard computation. First, for the blowup b, with center of codimension two, we have KC=b*KP3×Hilb2m+1(P3)+Eand for the product KP3×Hilb2m+1(P3)=pr1*KP3+pr2*KHilb2m+1(P3).

Now KP3=4H and for the projective bundle Hilb2m+1(P3)P(E) we have KP(E)=π*KPˇ3+c1(Ωπ1).

Again KPˇ3=4H and the short exact sequence 0Ωπ1π*(E)OP(E)(1)OP(E)0

gives c1(Ωπ1)=c1(π*(E)OP(E)(1))=π*c1(E)+6c1(OE(1)=π*c1(E)6A.

Putting this together, the stated expression for KC follows once we have established that c1(E)=4H.

Recall that E=pr2*(pr1*OP3(2)|I). We compute its first Chern class by brute force: apply Grothendieck–Riemann–Roch to pr2:P3×Pˇ3Pˇ3. Note that all higher direct images vanish, since Hp(V,OV(2))=0 for all VPˇ3 and p > 0. Thus by Grothendieck–Riemann–Roch the class c1(pr2*(pr1*OP3(2)|I))is the push forward in the sense of the Chow ring of the degree 4 homogeneous part of ch(pr1*OP3(2)|I)pr1*(td(P3)).

We have (5.1) td(P3)=(H1eH)4=1+2H+116H2+H3.(5.1)

Moreover IP3×Pˇ3 is a divisor of bidegree (1, 1), so there is a short exact sequence 0pr1*OP3(1)pr2*OPˇ3(1)OP3×Pˇ3OI0

from which we see (suppressing the explicit pullbacks pri* of cycles in the notation) (5.2) ch(pr1*OP3(2)|I))=exp(2H)(1exp(H)exp(H)).(5.2)

Now multiply together (5.1) and (5.2) and observe that the H3H-coefficient is 4. Since the push forward pr2* of any degree 4 monomial HkH4k equals H if k = 3 and 0 otherwise, this shows that c1(E)=4H. □

5.3 Basis for 1-cycles

We will need a few more effective curves, as before written as families {Yt}:

α: fix a conic C and a line L. Let the point Pt vary along L and let Yt=C{Pt}.

β: fix a quadric surface QP3, a line L and a point P. Let Vt run through the pencil of planes containing L and let Ct=QVt. Then take Yt=Ct{P}.

γ: fix a plane V and a point P. Let CtV run through a pencil of conics and let Yt=Ct{P}.

As before all choices are general, so that in the definition of α, the line L is disjoint from C, etc.

Lemma 5.5.

The dual basis to (H,H,A,E) is (α,β,γ,δ).

Proof.

We need to compute all the intersection numbers and verify that we get 0 or 1 as appropriate. Here it is sometimes useful to explicitly write out the pullbacks to C, e.g. writing b*(pr1*(H)) rather than H. We view α,β,γ,δ not just as equivalence classes, but as the effective curves defined above. Only the intersection numbers involving β require some real work, and we will save this for last.

Intersections with α: Since pr1*(b*(α))) is the line LP3 defining α we have b*(pr1*(H))·α=H·L=1. Similarly pr2*(b*(α))=0 shows that the intersections with H and A vanish. Finally α has no elements with embedded points, so is disjoint from E.

Intersections with γ: We have A·γ=1 because γ is a line in a fiber of the projective bundle π, whereas A restricts to a hyperplane in every fiber. The remaining intersection numbers vanish as we can pick disjoint effective representatives.

Intersections with δ: We have E·δ=1 as δ is a fiber of the blowup b and E is the exceptional divisor. The remaining divisors H, H, and A are all pullbacks, i.e. of the form b*(?) and then b*(?)·δ=(?)·b*(δ)=0.

Intersections with β: We can choose β to be disjoint from H and E. Moreover π*(pr1*(b*(β))) is the line LˇPˇ3 dual to the line L defining β. This gives b*(pr1*(π*(H)))·β=H·Lˇ=1.

It remains to verify A·β=0. The definition of β can be understood as follows: choose a general section OP3σOP3(2)

and apply pr2*(pr1*()|I) to obtain a homomorphism (5.3) OPˇ3E(5.3) whose fiber over VPˇ3 is exactly the restriction of σ to V. This is nowhere zero, so (5.3) is a rank 1 subbundle and it defines a section sQ:Pˇ3P(E)Hilb2m+1(P3) with sQ*(OP(E)(1))OPˇ3 or in terms of divisors sQ*(A)=0. If we let QP3 be the quadric defined by σ then sQ(V)=QV. Thus pr2*(b*(β)))=sQ*(Lˇ) where Lˇ is the dual to the line L defining β. This gives b*(pr2*(A))·β=A·pr2*(b*(β))=A·sQ*(Lˇ)=sQ*(A)·Lˇ=0.

We also define the following three effective divisors, phrased as a condition on YC:

D: all Y whose conic part C intersects a fixed line MP3.

D: all Y such that the line through P and a fixed point P0P3 intersects the conic part C.

E: all planar Y (as before).

Since D is defined by a condition on C only, it is the preimage by pr2°b (see Diagram Citation4.Citation1) of the similarly defined divisor in Hilb2m+1(P3). Moreover D and E are the strict transforms by b of the similarly defined divisors on P3×Hilb2m+1(P3).

We will need to control elements of D with an embedded point.

Lemma 5.6.

Fix P0 so that D is defined as an effective divisor. Choose a plane V not containing P0, a possibly degenerate conic CV and a point PC. Then there is a unique YD with conic part C and an embedded point at P. More precisely:

  1. If C is nonsingular at P then the embedded point structure is uniquely determined by the normal direction given by the line through P0 and P.

  2. If C is a pair of lines intersecting at P or a double line, then the embedded point is the spatial one, i.e. the scheme theoretic union of C and the first order infinitesimal neighborhood of P in P3.

Proof.

Let Q be the cone over C with vertex P0. This is a quadratic cone in the usual sense when C is nonsingular, otherwise Q is either a pair of planes or a double plane. A disjoint union C{P} with PP0 is clearly in D if and only if it is a subscheme of Q.

On the one hand this shows that the subschemes Y listed in (1) and (2) are indeed in D, since they are obtained from C{P} by letting P approach P along the line joining P0 and P.

On the other hand it follows that if YD then YQ, since the latter is a closed condition on Y. In case (1) Q is nonsingular at P and so there is a unique embedded point structure at PC which is contained in Q. In case (2) the following explicit computation gives the result: suppose in local affine coordinates that C is the pair of lines V(xy, z), the “vertex” P0 is on the z-axis and P is the origin. Then Q is the pair of planes V(xy). Any C with an embedded point at P has ideal of the form (xy,z)(x,y,z)+(sxy+tz)for (s:t)P1. This contains the defining equation xy of Q if and only if t = 0, which defines the spatial embedded point. The case where C is double line V(x2,z) is similar. □

Lemma 5.7.

We have D=2H+A,D=2H+2H+AE,E=H+HE.

Proof.

The last equality was essentially established in the proof of Proposition 4.3: it follows from the observations (1) E is the strict transform of b(E), and (2) the latter is the pullback of the incidence variety IP3×Pˇ3 which is linearly equivalent to H+H.

The remaining two identities are verified by computing the intersection numbers with the curves in the basis from Lemma 5.5. All curves and divisors involved are concretely defined and it is easy to find and count the intersections directly. Some care is needed to rule out intersection multiplicities, and we often find it most efficient to resort to a computation in local coordinates. We limit ourselves to writing out only two cases.

The case D·β=2: As we noted D is really a divisor on Hilb2m+1(P3) and so we shall write it here as b*(pr2*(D)). Then DHilb2m+1(P3) consists of all conics intersecting a fixed line M. We have b*(pr2*(D))·β=D·pr2*(b*(β))and pr2*(b*(β)) is the family of conics Ct=VtQ where Q is a fixed quadric surface and Vt runs through the pencil of planes containing a fixed line L. For general choices MQ consists of two points, and each point spans together with L a plane. This yields exactly two planes V0 and V1 in the pencil for which C0=V0Q and C1=V1Q intersects M. It remains to rule out multiplicities.

In the local coordinates in Remark 4.1 let M=V(x0,x1). Then the intersection between M and the plane x3=c0x0+c1x1+c2x2 is the point (0:0:1:c2). Now D is the condition that this point is on C, i.e. it satisfies Equationequation (4.3); this gives that D is s22=0. On the other hand, pr2*(b*(β)) is a one parameter family in which ci and sij are functions of degree at most 2 in the parameter. To stay concrete, let Q be ixi2=0 and let Vt be x3=tx2. Substitute x3 = tx2 in the equation for Q to find Ct=QVt. This gives in particular s22=1+t2 and so the intersection with D is indeed two distinct points, each of multiplicity 1.

The case D·δ=1: This is essentially Lemma 5.6, but to ascertain there is no intersection multiplicity to account for we argue differently. D is the strict transform of the divisor b(D)P3×Hilb2m+1(P3), which contains the center of the blowup. Since b(D) is nonsingular (pick P0=(0:0:0:1) in the definition of D, then in the local coordinates of Remark 4.1 it is simply given by the Equationequation (4.3)) we have D=b*(b(D))E. Thus D·δ=b(D)·b*(δ)E·δ=0(1)and we are done.

The remaining cases are either similar to these or easier. □

5.4 Nef and Mori cones

It is clear that H, H and D are base point free, hence nef. For instance, consider D: given YC, choose a line MP3 disjoint from YP3. This defines an effective representative for D not containing Y.

Lemma 5.8.

The divisor D+H is nef.

Proof.

We begin by narrowing down the base locus of D. First consider an element YC without embedded point, that is a disjoint union Y=C{P}. Then choose P0 such that the line through P0 and P is disjoint from C. This defines a representative for D not containing Y, so Y is not in the base locus.

Next let Y be a conic C with an embedded point at a point PC where C is nonsingular. The tangent to C at P together with the normal direction given by the embedded point determines a plane. Pick P0 such that the line through P and P0 defines a normal direction to C which is distinct from that defined by the embedded point. This determines a representative for D which by Lemma 5.6(i) does not contain Y, so Y is not in the base locus.

The remaining possibility is that Y is either a pair of intersecting lines with an embedded point at the singularity, or a double line with an embedded point. Pick a representative for D by choosing P0 outside the plane containing the degenerate conic. If the embedded point is not spatial, then Lemma 5.6(ii) shows that Y is not in D. So Y is not in the base locus unless the embedded point is spatial.

Thus let BC be the locus of intersecting lines with a spatial embedded point at the origin, together with double lines with a spatial embedded point. By the above B contains the base locus of D, so if TC is an irreducible curve not contained in B then (D+H)·T=D·T+H·T0as both terms are nonnegative. If on the other hand TB we observe that T·E=0: in fact B and E are disjoint, since every element in B has a spatial embedded point, whereas all elements in E are planar. By the relations in Lemma 5.7 D+H=H+D+E and so, using that H and D are nef, (D+H)·T=(H+D+E)·T=H·T+D·T0+E·T00.

Lemma 5.9.

The dual basis to (H,H,D,D+H) is (ϵ,η,ζ,δ).

Proof.

Lemma 5.5 together with the relations in Lemma 5.7 implies that (D+H)·δ=1 and the other three intersection numbers with δ vanish.

Of the remaining intersection numbers only those involving η requires some care and we shall write out only those.

A representative for η is obtained by fixing a line L and a point PL and letting the plane Vt vary in the pencil of planes containing Vt. Then Ct is the double L inside Vt and Yt is Ct together with an embedded spatial point at P. Then:

  • Intersecting with H imposes the condition that P is contained in a fixed but arbitrary plane, but P is fixed, so H·η=0.

  • Intersecting with H imposes the condition that Vt contains a fixed but arbitrary point P0, this gives H·η=1. (In fact this can be identified with the intersection number H·Lˇ=1 in Pˇ3, where Lˇ is the dual line to L, so there is no subtlety regarding transversality of the intersection.)

  • Intersecting with D imposes the condition that Ct intersects a fixed but arbitrary line M, but Ct has fixed support L, so D·η=0.

As η is contained in the base locus of D we cannot find D·η directly. As in the proof of Lemma 5.8 we instead rewrite D+H as H + D + E and take advantage of η being disjoint from E. This gives (D+H)·η=(H+D+E)·η=0.

We wish to point out that the computation in the very last paragraph, showing D·η=1, is what made us realize that the addition of H is necessary to produce a nef divisor.

Proof of Theorem 5.1.

The four divisors in Lemma 5.9 are nef (the first three are base point free, and the fourth is treated in Lemma 5.8) and the four curves are effective by definition. Hence they span the nef and Mori cones of C, respectively. Finally by Lemmas 5.4 and 5.7 we have K=2H+5H5D(D+H)so in view of the dual bases in Lemma 5.9 we read off that K is negative on ϵ, ζ, δ and positive on η. □

Additional information

Funding

This work is supported by the Research Council of Norway under Grant No. 230986, and is part of the Ph.D. thesis [Citation1] defended by the first author at the university of Stavanger on the 29th of March 2023. The second author was also supported by the Swedish Research Council under Grant No. 2016-06596 while the author was in residence at Institut Mittag–Leffler in Djursholm, Sweden during the fall of 2021.

References

  • Alaoui Soulimani, S. (2023). Bridgeland stability conditions and the Hilbert scheme of skew lnes in projective space. PhD thesis, University of Stavanger. Available at https://hdl.handle.net/11250/3058550.
  • Artin, M. (1970). Algebraization of formal moduli: II. Existence of modifications. Ann. Math. 91(1):88–135. DOI: 10.2307/1970602.
  • Bayer, A., Macrì, E. (2011). The space of stability conditions on the local projective plane. Duke Math. J. 160(2):263–322. DOI: 10.1215/00127094-1444249.
  • Bayer, A., Macrì, E., Stellari, P. (2016). The space of stability conditions on Abelian threefolds, and on some Calabi-Yau threefolds. Invent. Math. 206(3):869–933. DOI: 10.1007/s00222-016-0665-5.
  • Bayer, A., Macrì, E., Toda, Y. (2014). Bridgeland stability conditions on threefolds I: Bogomolov-Gieseker type inequalities. J. Algebraic Geom. 23(1):117–163. DOI: 10.1090/S1056-3911-2013-00617-7.
  • Bridgeland, T. (2007). Stability conditions on triangulated categories. Ann. Math. 166(2):317–345. DOI: 10.4007/annals.2007.166.317.
  • Bridgeland, T. (2008). Stability conditions on K3 surfaces. Duke Math. J. 141(2):241–291. DOI: 10.1215/S0012-7094-08-14122-5.
  • Chen, D., Coskun, I., Nollet, S. (2011). Hilbert scheme of a pair of codimension two linear subspaces. Commun. Algebra 39(8):3021–3043. DOI: 10.1080/00927872.2010.498396.
  • Chen, D., Nollet, S. (2012). Detaching embedded points. Algebra Number Theory 6(4):731–756. DOI: 10.2140/ant.2012.6.731.
  • Fujiki, A., Nakano, S. (1971). Supplement to “On the Inverse of Monoidal Transformation”. Publ. Res. Inst. Math. Sci. (Kyoto) 7(3):637–644. DOI: 10.2977/prims/1195193401.
  • Gallardo, P., Lozano Huerta, C., Schmidt, B. (2018). Families of elliptic curves in P3 and Bridgeland stability. Michigan Math. J. 67(4):787–813.
  • Lascu, A. T. (1969). Sous-variétés régulièrement contractibles d’une variété algébrique. Annali della Scuola Normale Superiore di Pisa - Classe di Scienze 23(4):675–695.
  • Lee, Y. H. A. (2000). The Hilbert scheme of curves in P3. Bachelor’s thesis, Harvard university.
  • Macrì, E. (2014). A generalized Bogomolov-Gieseker inequality for the three-dimensional projective space. Algebra Number Theory 8(1):173–190. DOI: 10.2140/ant.2014.8.173.
  • Macrì, E., Schmidt, B. (2017). Lectures on Bridgeland stability. In: Moduli of Curves, volume 21 of Lecture Notes of the Unione Matematica Italiana. Cham: Springer, pp. 139–211.
  • Moishezon, B. G. (1967). On n-dimensional compact complex manifolds having n algebraically independent meromorphic functions. Amer. Math. Soc. 63:51–177.
  • Nakano, S. (1970). On the Inverse of Monoidal Transformation. Publ. Res. Inst. Math. Sci. (Kyoto) 6(3):483–502. DOI: 10.2977/prims/1195193917.
  • Piyaratne, D., Toda, Y. (2019). Moduli of Bridgeland semistable objects on 3-folds and Donaldson-Thomas invariants. J. Reine Angew. Math. (Crelle’s Journal) 747:175–219. DOI: 10.1515/crelle-2016-0006.
  • Rezaee, F. (2021). Geometry of canonical genus four curves. Preprint arXiv:2107.14213 [math.AG].
  • Schmidt, B. (2020). Bridgeland stability on threefolds: some wall crossings. J. Algebraic Geom. 29(2):247–283. DOI: 10.1090/jag/752.
  • The Stacks project authors. (2023). The stacks project. https://stacks.math.columbia.edu, 2023.
  • Xia, B. (2018). Hilbert scheme of twisted cubics as a simple wall-crossing. Trans. Amer. Math. Soc. 370(8):5535–5559. DOI: 10.1090/tran/7150.